+ All Categories
Home > Documents > Yale University Department of Computer Sciencecs.yale.edu/publications/techreports/tr1389.pdf ·...

Yale University Department of Computer Sciencecs.yale.edu/publications/techreports/tr1389.pdf ·...

Date post: 21-Oct-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
43
Yale University Department of Computer Science Cryo-EM Structure Determination through Eigenvectors of Sparse Matrices Ronald. R. Coifman Yoel Shkolnisky Fred J. Sigworth Amit Singer YALEU/DCS/TR-1389
Transcript
  • Yale University

    Department of Computer Science

    Cryo-EM Structure Determination through

    Eigenvectors of Sparse Matrices

    Ronald. R. Coifman Yoel ShkolniskyFred J. Sigworth Amit Singer

    YALEU/DCS/TR-1389

  • Cryo-EM Structure Determination through

    Eigenvectors of Sparse Matrices

    Ronald. R. Coifman∗, Yoel Shkolnisky∗, Fred J. Sigworth†

    and Amit Singer∗

    Abstract

    Recovering the three-dimensional structure of proteins is impor-

    tant for understanding their functionality. We describe a spectral

    graph algorithm for reconstructing the three-dimensional structure of

    molecules from their cryo-electron microscopy images taken at random

    unknown orientations. The key idea of the algorithm is designing a

    sparse operator defined on the projection images, whose eigenvectors

    reveal their orientations.

    The special geometry of the problem rendered by the Fourier projection-

    slice theorem is incorporated into the construction of a weighted graph

    whose vertices are the radial Fourier lines and whose edges are linked

    with the common line property. The radial lines are associated with

    points on the sphere and are networked through spider like connec-

    tions. The graph organizes the radial lines on the sphere in a global

    manner that reveals the projection directions. This organization is de-

    rived from a global computation of a few eigenvectors of the graph’s

    adjacency matrix. Once the directions are obtained, the molecule can

    be reconstructed using classical tomography methods.

    ∗Department of Mathematics, Program in Applied Mathematics, Yale Univer-sity, 10 Hillhouse Ave. PO Box 208283, New Haven, CT 06520-8283 USA.{[email protected], [email protected], [email protected]}

    †Department of Cellular and Molecular Physiology, Yale University School of Medicine,333 Cedar Street, New Haven, CT 06520 USA. [email protected]

    2

  • The presented algorithm is direct (as opposed to iterative refine-

    ment schemes), does not require any prior model for the reconstructed

    object, and shown to have favorable computational and numerical

    properties. Moreover, the algorithm does not impose any assump-

    tion on the distribution of the projection orientations. Physically, this

    means that the algorithm successfully reconstructs molecules that have

    unknown spatial preference.

    We also introduce extensions of the algorithm, based on the spec-

    tral properties of the operator, which significantly improve its ap-

    plicability to realistic data sets. These extensions include: particle

    selection, to filter corrupted projections; center determination to es-

    timate the relative shift of each projection; and, a method to resolve

    the heterogeneity problem, in cases where a mix of different molecules

    is being imaged concurrently.

    1 Introduction

    “Three dimensional electron microscopy” [1] is the name commonly given

    to methods in which the 3D structures of macromolecular complexes are

    obtained from sets of images taken by an electron microscope. The most

    widespread and general of these methods is single-particle reconstruction

    (SPR). In SPR the 3D structure is determined from images of randomly

    oriented and positioned identical macromolecular “particles”, typically com-

    plexes 500 kDa or larger in size. The SPR method has been applied to

    images of negatively stained specimens, and to images obtained from frozen-

    hydrated, unstained specimens [2]. In the latter technique, called cryo-

    electron microscopy (cryo-EM) the sample of macromolecules is rapidly frozen

    in a thin (∼ 100 nm) layer of vitreous ice, and maintained at liquid nitrogen

    temperature throughout the imaging process.

    SPR from cryo-EM images is of particular interest because it promises

    to be an entirely general technique. It does not require crystallization or

    3

  • other special preparation of the complexes to be imaged, and in the future

    it is likely to reach sufficient resolution (∼ 0.4 nm) to allow the polypeptide

    chain to be traced and residues identified in protein molecules [3]. Even

    at the present best resolutions of 0.9–0.6 nm, many important features of

    protein molecules can be determined [4].

    Much progress has been made in algorithms that, given a starting 3D

    structure, are able to refine that structure on the basis of a set of negative-

    stain or cryo-EM images, which are taken to be projections of the 3D object.

    Data sets typically range from 104 to 105 particle images, and refinements

    require tens to thousands of CPU-hours. As the starting point for the refine-

    ment process, however, some sort of ab initio estimate of the 3D structure

    must be made. Present algorithms are based on the “Angular Reconstitu-

    tion” method of van Heel [5] in which a coordinate system is established from

    three projections, and the orientation of the particle giving rise to each image

    is deduced from common lines among the images.

    A serious limitation and an active area of research in SPR image process-

    ing is the problem of heterogeneity [6, 7]. Often a set of images is obtained

    from a mixture of particles of two or more different kinds or different confor-

    mations. Because traditional SPR methods assume identical particles, they

    fail to distinguish the different particle species. Extremely desirable would

    be the ability to establish two or more ab initio reconstructions from a single

    set of images.

    We propose Globally Consistent Angular Reconstitution (GCAR), a recon-

    struction algorithm that does not assume any ab initio model and establishes

    a globally consistent coordinate system from all projections. The special ge-

    ometry of the problem rendered by the Fourier projection-slice theorem [8]

    is incorporated by GCAR into a weighted directed graph whose vertices are

    the radial Fourier lines and whose edges are linked using the common line

    property. Radial lines are viewed as points on the sphere and are networked

    through spider-like connections. The graph organizes the radial lines on the

    4

  • sphere in a global manner and reveals the projection directions. Such an

    organization is derived from a global computation of the eigenvectors of the

    sparse graph matrix. This global averaging property makes GCAR robust

    to both noise and false detections of common lines. The global organization

    of the common lines also promises to provide a solution to the problems of

    center determination and heterogeneity. GCAR is extremely fast because it

    requires only the computation of a few eigenvectors of a sparse matrix.

    Once the directions are revealed by the eigenvectors, the reconstruction is

    performed using the combination of classical tomography methods together

    with more recent non-uniform FFT techniques [9, 10].

    We have successfully applied similar graph-based approaches to recon-

    struct two-dimensional structures, such as the Shepp-Logan phantom, from

    noisy one-dimensional projections taken at random directions [11]. Many of

    the recent and successful algorithms for nonlinear dimensionality reduction of

    high-dimensional data, such as locally linear embedding (LLE) [12], Hessian

    LLE [13], Laplacian eigenmap [14] and diffusion maps [15] involve the com-

    putation of eigenvectors of data-dependent sparse kernel matrices. However,

    such algorithms fail to solve the cryo-EM problem, because the reduced co-

    ordinate system that each of them obtains does not agree with the projection

    directions. On the other hand, GCAR finds the desired coordinate system

    of projection images, because it is tailored to the geometry of the problem

    through the Fourier projection-slice theorem.

    The organization of this report is as follows. In Section 2 we introduce the

    cryo-electron microscopy problem, as well as the related mathematical back-

    ground. We introduce the GCAR operator, which reveals the orientations of

    the projections, and analyze its properties in Section 3. The algorithm for

    recovering the projection orientations is summarized in Section 4, together

    with a few technical implementation details. Examples of applying this algo-

    rithm to two phantoms are given in Section 5. In Section 6 we present several

    useful extensions of the algorithm. In particular, we present how to deter-

    5

  • mine the center of each projection by utilizing the common-line property in

    Section 6.1, how to solve the heterogeneity problem, where projections of two

    types of molecules are mixed together, in Section 6.2, and how to exploit the

    properties of the GCAR graph to filter out “bad” projections in Section 6.3.

    We conclude in Section 7 with a summary and a description of future work.

    2 Problem Setup

    2.1 The physical setting

    The cryo-EM reconstruction problem is to find the three-dimensional struc-

    ture of a molecule given samples of its two-dimensional projection images at

    unknown random directions. The intensity of pixels in a given projection

    image corresponds to line integrals of the electric potential induced by the

    molecule along the path of the electrons. The highly intense electron beam

    destroys the molecule and it is therefore impractical to take projection im-

    ages of the same molecule at known different directions, as in the case of

    classical computerized tomography (CT). In other words, a single molecule

    can be imaged only once. All molecules are assumed to have the exact same

    structure; they differ only by their spatial orientation. Thus, every image is

    a projection of the same molecule but at an unknown random orientation.

    The locations of the microscope (source) and the camera/film (detectors)

    are fixed, while different images correspond to different spatial rotations of

    the molecule. Every image is thus associated with an element of the rotation

    group SO(3). If the electric potential of the molecule is φ(r) in some fixed

    ‘laboratory’ coordinate system r = (x, y, z), then rotating the molecule by

    g ∈ SO(3) results in the potential φg(r) = φ(g−1r). The projection image

    Pg(x, y) is obtained by integrating φg(r) along the z-direction (the source-

    detector direction)

    Pg(x, y) =

    ∫ ∞

    −∞

    φg(x, y, z) dz.

    6

  • Projection

    Molecule

    Electronsource

    g∈SO(3)

    Figure 2.1: Schematic drawing of projection.

    The projection operator is also known as the X-ray transform [8]. Figure 2.1

    is a schematic illustration of the cryo-EM setting.

    The projection image is a digital picture given as an p× p grid of pixels

    Pg = {Pg(xi, yj)}pi,j=1 ∈ R

    p2 , where p is determined by the characteristics of

    the imaging setup.

    The cryo-EM problem is stated as follows: find φ(x, y, z) given a collection

    of K projections {Pgk}Kk=1, where gk are unknown rotations. If the rotations

    {gk}Kk=1 were known then the reconstruction of φ(x, y, z) could be performed

    by classical tomography methods. Therefore, the cryo-EM problem is re-

    duced to estimating the rotations {gk}Kk=1 given the data set {Pgk}

    Kk=1.

    For convenience, we adopt the following equivalent point of view. Instead

    of having the microscope fixed and the molecule oriented randomly in space,

    we think of the molecule as being fixed, and the microscope being the one that

    is randomly rotated in space. The orientation of the microscope is described

    by the beaming direction θ ∈ S2 (axis of rotation) and the in-plane rotation

    7

  • angle α ∈ [0, 2π) of the camera.

    2.2 Fourier projection-slice theorem

    The two-dimensional Fourier transform of a projection image P (x, y) is given

    by the double integral

    P̂ (ω) =1

    (2π)2

    θ⊥e−ir·ωP (r) dr, (1)

    where θ is the projection direction, that is, a normal to the image plane. The

    three-dimensional Fourier transform of the molecule is given by the triple

    integral

    φ̂(ξ) =1

    (2π)3

    R3

    e−ir·ξφ(r) dr. (2)

    One of the cornerstones of tomography is the Fourier projection-slice theorem

    stating that the two-dimensional Fourier transform of a projection image is

    a planar slice θ⊥ of the three-dimensional Fourier transform of the molecule

    (see, e.g., [8, p. 11])

    P̂ (η) = 2πφ̂(η), η ∈ θ⊥. (3)

    Figure 2.2 depicts the geometry induced by the Fourier projection-slice the-

    orem. The sphere represents the 3D Fourier space of the molecule, and the

    planes are three particular two-dimensional slices of it. As the 2D Fourier

    transform of a projection image corresponds to a slice of the 3D Fourier

    space, any two slices share a common line, i.e., the intersection line of the

    two planes. Figure 2.2 shows three particular slices that happen to share the

    same common line, marked in dashed black. Note that the effect of rotating

    the camera (changing α) while fixing the beaming direction θ is an in-plane

    rotation of the slice without changing its position.

    The Fourier transform of each projection image can be considered in polar

    coordinates as a set of planar radial lines in two-dimensional Fourier space.

    As a result of the Fourier projection-slice theorem, every planar radial line

    8

  • Figure 2.2: Two-dimensional slices of the three-dimensional Fourier space.

    is also a radial line in the three-dimensional frequency ball of Fig. 2.2, and

    there is a one-to-one correspondence between those Fourier radial lines and

    points on the unit sphere S2. The radial lines of a single projection image

    correspond to a great circle (a geodesic circle) on S2. For example, the radial

    lines of the green slice in Fig. 2.2 form the equator. Thus, to every projection

    image Pgk there corresponds a unique great circle Ck over S2, and the common

    line property is restated as follows: any two different geodesic circles over S2

    intersect at exactly two antipodal points.

    If the projection directions (the gk’s or the (θk, αk)’s) were known, then

    the 2D polar Fourier transform of the projection images would have resulted

    in the values of φ̂(ξ) on different 3D radial lines, as stated by Eq. 3. Fourier

    inversion of φ̂(ξ) in a frequency ball of radius B (|ξ| < B, where B is de-

    termined by the sampling resolution) would have then reconstructed a band

    limited approximation of φ. However, in the cryo-EM problem the slices are

    unorganized. Neither their directions θ nor their in-plane rotations α are

    known.

    9

  • 2.3 Correspondence between Fourier rays and S2

    We next explain the discretization of Fourier space and derive a mapping

    between the discretized Fourier space and points on the unit sphere S2. Such

    a mapping would allow us to proceed by exploiting the simple geometry of

    S2.

    Let P1(x, y), . . . , PK(x, y) be K projection images. Upon writing the

    Fourier transform in Eq. 1 in polar coordinates, we obtain

    P̂k(ρ, ω) =1

    (2π)2

    ∫ ∫

    Pk(x, y)e−i(xρ cos ω+yρ sinω) dx dy, k = 1, . . . , K. (4)

    For digital implementations we discretize ρ and ω, and compute Eq. 4 using a

    nonequally-spaced FFT [9]. We denote by L the angular discretization of ω,

    and sample ρ in n equally-spaced points. That is, we split each transformed

    projection into L radial lines Λk,0, . . . ,Λk,L−1 ∈ Rn, each represented by a

    set of n equispaced points

    Λk,l =(

    P̂k(B/n, 2πl/L), P̂k(2B/n, 2πl/L), . . . , P̂k(B, 2πl/L))

    ∈ Rn, (5)

    for 1 ≤ k ≤ K, 0 ≤ l ≤ L− 1, where B is the band limit. Note that the DC

    term (ρ = 0 in Eq. 4) is shared by all lines independently of the image and

    can therefore be ignored.

    Let

    Λ(β) =(

    2πφ̂(βB/n), 2πφ̂(2βB/n), . . . , 2πφ̂(nβB/n))

    ∈ Rn (6)

    be n samples from a ray through the origin in 3D Fourier space, in a direction

    given by the unit vector β ∈ S2. This defines a map Λ : S2 → Rn that maps

    each unit vector in R3 to n samples of the Fourier ray in that direction.

    According to the Fourier projection-slice theorem, the radial lines Λk,l are

    the evaluations of the function Λ(β) at the unknown points βk,l ∈ S2. The

    function Λ(β) is unknown, because so is φ̂. Our goal is to find the sources

    10

  • βk,l of the overall KL radial lines. The L sources {βk,l}L−1l=0 (k fixed) are

    equidistant points on the great circle Ck ⊂ S2.

    3 Orientation Revealing Operator

    In this section we introduce the GCAR operator, whose eigenvectors reveal

    the orientation of each projection. The GCAR operator is a graph with every

    node representing a ray in Fourier space, and whose edges are determined by

    the common-line property. The formal construction of this graph is presented

    in Section 3.1. The normalized adjacency matrix of the graph can be viewed

    as an averaging operator for functions defined on the nodes of the graph,

    as explained in Section 3.2. Analyzing the eigenvectors of this operator in

    Section 3.3 shows that they encode the projection orientations. We conclude

    by showing in Section 3.4 that the eigenvectors of the GCAR matrix are

    intimately related to the spherical harmonics.

    3.1 GCAR graph

    We assume as before that we are given K projection images. Let Λk,l, k =

    1, . . . , K, l = 1, . . . , L be KL Fourier rays computed from the K projection

    images. We think of the radial lines as vertices of a graph, where each radial

    line Λk,l and its source βk,l are identified with the vertex indexed by the pair

    (k, l). In other words, the set of vertices V of the directed graph G = (V,E)

    is

    V = {(k, l) : 1 ≤ k ≤ K, 0 ≤ l ≤ L− 1},

    where the number of vertices is |V | = KL. The set E ⊂ V × V of directed

    edges (or arrows)

    E = {((k1, l1), (k2, l2)) : (k1, l1) points to (k2, l2)}

    11

  • will be defined shortly. The graphG is represented using a (sparse) adjacency

    matrix W of size KL×KL

    W(k1,l1),(k2,l2) =

    {

    1 if ((k1, l1), (k2, l2)) ∈ E

    0 if ((k1, l1), (k2, l2)) 6∈ E.

    Each row of W corresponds to one radial line in Fourier space Λk,l, or al-

    ternatively to a point βk,l on S2. The nonzero elements in that row, when

    viewed on S2, constitute a spider-like neighborhood, as will be explained

    shortly.

    We now define the set E. Consider a specific vertex (k1, l1), that is, the

    radial line Λk1,l1. Identify its 2J + 1 neighboring radial lines from its own

    transformed projection P̂gk1 (indicated in red in the top panel of Fig. 3.1)

    and set

    ((k1, l1), (k1, l1 + l)) ∈ E for − J ≤ l ≤ J, (7)

    where addition is modulo L. We link every radial line with J neighboring

    radial lines in each direction, clockwise and counterclockwise.

    We next determine the remaining nonzero entries in row (k1, l1) of W by

    using the common-line property. Whenever we identify Λk1,l1 ≈ Λk2,l2 as the

    common line between projection images k1 and k2, we add to E the following

    links

    ((k1, l1), (k2, l2 + l)) ∈ E for − J ≤ l ≤ J. (8)

    The antipodal rays Λk1,l1+L/2 and Λk2,l2+L/2 are common radial lines as well

    and are linked together in the same manner by setting

    ((k1, l1 + L/2), (k2, l2 + L/2 + l)) ∈ E for − J ≤ l ≤ J. (9)

    In general, the adjacency matrix W is nonsymmetric: ((k1, l1), (k2, l2)) ∈

    E (for k1 6= k2) reflects the fact that the circle Ck2 containing (k2, l2) passes

    nearby the point (k1, l1). However, Ck1 does not necessarily passes nearby

    12

  • Projection Fourier transform 3D Fourierspace

    3D Fourierspace

    Fourier transformProjection

    Pgk1 P̂gk1

    (k1,l1)

    Λk2,l2=Λk1,l1Λk2,l2=Λk1,l1

    Pgk2 P̂gk2

    Figure 3.1: Fourier projection-slice theorem.

    (k2, l2) so ((k2, l2), (k1, l1)) 6∈ E. Symmetry occurs only within the same

    circle, that is, ((k, l1), (k, l2)) ∈ E ⇐⇒ ((k, l2), (k, l1)) ∈ E, which happens

    whenever |l1 − l2| ≤ J . Although the size of W is KL × KL, by choosing

    J ≪ L we force it to be a sparse matrix. Its number of nonzero entries is

    only

    |E| = (2J + 1)KL+ 2K(K − 1)(2J + 1). (10)

    The first summand corresponds to the 2J + 1 edges in Eq. 7 for each of the

    KL vertices. The second summand corresponds to edges between different

    images (Eq. 8). Any two circles intersect at exactly two antipodal points,

    so there are 2(

    K2

    )

    = K(K − 1) meeting points. Every meeting point, that

    is, every common-line Λk1,l1 = Λk2,l2 contributes 2J + 1 nonzero elements in

    row (k1, l1) of W , and 2J + 1 nonzero elements in row (k2, l2). In practice,

    the number of edges in E is smaller than the number in Eq. 10, as explained

    in Section 4.

    If we take row (k1, l1) from W and plot the points βk,l ∈ S2 for which

    13

  • (a) One row of W – one spider (b) Two rows of W – two spiders

    Figure 3.2: Mapping the nonzero entries of W to S2.

    W (k1,l1),(k,l) = 1, we get a picture that looks like Fig. 3.2a. In light of

    Fig. 3.2a, we refer to each row of W as the “spider that corresponds to the

    point (k1, l1)”. The point βk1,l1 is the head (center) of the spider. Sources

    βk,l that correspond to the same k (come from the same projection image)

    are marked with the same color. Figure 3.2b shows two spiders on S2 for a

    certain randomized choice of circles with K = 200, L = 100, J = 10, from

    which we can see how different spiders interact. This interaction is essential

    for the global consistent assignment of coordinates explained below.

    Different spiders may have different number of legs, so row sums of W

    may be different. The outdegree dk,l of the (k, l)’th vertex is the sum of the

    corresponding row in W

    dk,l =∑

    (k′,l′)∈V

    W(k,l),(k′,l′) = |{(k′, l′) : ((k, l), (k′, l′)) ∈ E}|. (11)

    14

  • 3.2 Averaging operator

    We normalize the adjacency matrixW to have constant row sums by dividing

    it by a diagonal matrix D whose diagonal elements equal the outdegrees dk,l

    of the vertices (or equivalently the row sums of W )

    D(k,l),(k,l) = dk,l, (12)

    with dk,l given by Eq. 11. This normalization results in the operator

    A = D−1W . (13)

    The operator A : C|V | → C|V | takes any discrete complex valued function

    f : V → C (realized as a vector in CKL) and assigns the head of each spider

    the average of f over the entire spider

    (Af )(k1, l1) =1

    dk1,l1

    ((k1,l1),(k2,l2))∈E

    f(k2, l2).

    We therefore regard A as an averaging operator over C|V |.

    The matrix A is row stochastic (the row sums of A all equal 1), and

    therefore the constant function ψ0 = 1 (ψ0(v) = 1, ∀v ∈ V ) is an eigenvector

    with λ0 = 1: Aψ0 = ψ0. The remaining eigenvectors may be complex

    and come in conjugate pairs, because A is real but not symmetric: Aψ =

    λψ ⇐⇒ Aψ̄ = λ̄ψ̄. As of the spectrum of A, λ0 = 1 is the largest possible

    eigenvalue and the remaining eigenvalues reside inside the complex unit disk

    |λ| < 1.

    3.3 Coordinate eigenvectors

    The operator A has many interesting properties. For the cryo-EM problem,

    the most important property is that the coordinates of the sources βk,l are

    eigenvectors of the averaging operator A, sharing the same eigenvalue. Ex-

    15

  • plicitly, let x, y, z : R3 → R be the coordinate functions in R3. Then, the

    vectors x,y, z ∈ RKL defined by

    x = x(βk,l), y = y(βk,l), z = z(βk,l),

    k = 1, . . . , K, l = 1, . . . , L are eigenvectors of A.

    This remarkable fact is a consequence of the following observation: the

    center of mass of every spider is in the direction of the spider’s head, because

    any pair of opposite legs balance each other. For example, the center of mass

    of a spider whose head is located at the north pole lies just beneath it.

    We include the details of the rather technical proof. Suppose f(x, y, z) =

    a1x + a2y + a3z = a · β is a linear function, where a = (a1, a2, a3) and

    β = (x, y, z) ∈ S2. Consider a spider whose head is at the point β1 =

    (x1, y1, z1) ∈ S2, where the value of the function is f(x1, y1, z1) = a · β1.

    Let β2,β3 be two unit vectors that complete β1 into an orthonormal system

    of R3. In other words, the 3 × 3 matrix U whose columns are β1,β2,β3

    is orthogonal. We express any point β = (x, y, z) on the sphere as a linear

    combination β = x′β1+y′β2+z

    ′β3 = Uβ′, where β′ = (x′, y′, z′) is a rotated

    coordinate system. We apply a change of variable β → β′ in f to obtain

    the linear function f ′(x′, y′, z′) = f(x, y, z) = a · β = a · Uβ′ = a′ · β′,

    where a′ = UTa = (a′1, a′2, a

    ′3). The parameterization of a great circle going

    through β1 is

    cos θβ1 + sin θ cosϕ0β2 + sin θ sinϕ0β3,

    where θ ∈ (−π, π] and ϕ0 is a fixed parameter that determines the direction

    of the circle. On that circle, f is a function of the single parameter θ

    f(θ) = f ′(cos θ, sin θ cosϕ0, sin θ sinϕ0) = a′ · (cos θ, sin θ cosϕ0, sin θ sinϕ0).

    16

  • The average f̄ of f over the two discrete opposite legs of that circle is

    f̄(x1, y1, z1) =1

    2J + 1

    J∑

    l=−J

    f

    (

    2πl

    L

    )

    =a′

    2J + 1·

    J∑

    l=−J

    (cos2πl

    L, sin

    2πl

    Lcosϕ0, sin

    2πl

    Lsinϕ0)

    =

    [

    1

    2J + 1

    J∑

    l=−J

    cos2πl

    L

    ]

    a′ · (1, 0, 0),

    due to the linearity of the dot product and the fact that sin θ is an odd

    function. From

    a′ · (1, 0, 0) = UTa · (1, 0, 0) = a ·U(1, 0, 0) = a · β1 = f(x1, y1, z1),

    we conclude that

    f̄(x1, y1, z1) =

    [

    1

    2J + 1

    J∑

    l=−J

    cos2πl

    L

    ]

    f(x1, y1, z1) (14)

    holds for all (x1, y1, z1) and for any circle going through it. Therefore, linear

    functions are eigenvectors of the averaging operator A with eigenvalue λ =1

    2J+1

    ∑Jl=−J cos

    2πlL

    . This completes the proof.

    3.4 Spherical harmonics

    The eigenfunctions of the Laplacian operator on the sphere S2 are known to

    be the spherical harmonics Y ml [8, p.195](also known as the eigenstates of

    the angular momentum operator in quantum mechanics)

    ∆S2Yml = −l(l + 1)Y

    ml , l = 0, 1, 2, . . . , m = −l,−l + 1, . . . , l. (15)

    17

  • The (non-normalized) spherical harmonics are given in terms of the associ-

    ated Legendre polynomials of the zenith angle θ ∈ [0, π] and trigonometric

    polynomials of the azimuthal angle ϕ ∈ [0, 2π)

    Y 0l (θ, ϕ) = Pl(cos θ),

    Y ml (θ, ϕ) = P|m|l (cos θ) cosmϕ, 1 ≤ m ≤ l,

    Y −ml (θ, ϕ) = P|m|l (cos θ) sinmϕ, 1 ≤ m ≤ l,

    while the Laplacian is given by

    ∆S2 =1

    sin θ

    ∂θ

    (

    sin θ∂

    ∂θ

    )

    +1

    sin2 θ

    ∂2

    ∂ϕ2. (16)

    The eigenspaces are degenerated so that the eigenvalue l(l + 1) has multi-

    plicity 2l+1. Alternatively, the l’th eigenspace corresponds to homogeneous

    polynomials of degree l restricted to S2. In particular, the first three non-

    trivial spherical harmonics Y m1 share the same eigenvalue and are given by

    the three linear functions

    Y 11 = x, Y−11 = y, Y

    01 = z.

    The spherical harmonics Y ml are usually derived by separating variables

    in Eqs. 15–16. However, the fundamental reason for which the spherical har-

    monics are eigenfunctions of the Laplacian is that the latter commutes with

    rotations. In fact, the classical Funk-Hecke theorem (see, e.g., [8, p. 195])

    asserts that the spherical harmonics are the eigenfunctions of any integral op-

    erator K : L2(S2) → L2(S2) that commutes with rotations. Such operators

    are of the form

    (Kf)(β) =

    S2k(〈β,β′〉)f(β′) dSβ′,

    where k : [−1, 1] → R is the kernel function that depends only on the angle

    18

  • between β,β′ ∈ S2. For such integral operators we have

    KY ml = λlYml ,

    where the eigenvalues λl depend on the specific kernel function k(·) and are

    given by

    λl = 2π

    ∫ 1

    −1

    k(t)Pl(t) dt.

    For example, the spherical harmonics are the eigenfunctions of the operator

    that corresponds to averaging over spherical caps.

    The averaging operator A defined in Section 3.2 does not commute with

    rotations, because every spider has different number of legs that go in dif-

    ferent directions. The averaging operator commutes with rotations only in

    the limit of infinite number of projection images corresponding to a uniform

    distribution over SO(3) (the Haar measure). Although A does not commute

    with rotations and the Funk-Hecke theorem does not hold, the coordinate

    vectors x,y, z ∈ RKL span an eigenspace of A, due to the center of mass

    property. Figure 3.3 depicts the first 50 eigenvalues of the operator A con-

    structed using K = 200 random points on S2, L = 100 points on each

    geodesic circle, and J = 10 samples on each leg of the spider. This corre-

    sponds to using K = 200 projection images, L = 100 radial Fourier lines for

    each projection, and using J = 10 in Eqs. 7–9. The threefold multiplicity

    corresponding to the coordinate vectors is clearly seen in Fig. 3.3. Moreover,

    the observed numerical multiplicities of 1, 3, 5, 7 and 9 are explained by the

    spherical harmonics. The remaining eigenvalues seen in Fig. 3.3 correspond

    to clustering of the circles. By clustering we mean that each of the corre-

    sponding eigenvectors takes an almost constant value on one of the circles

    and practically vanishes on all other circles.

    19

  • 0 5 10 15 20 25 30 35 40 450

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    Figure 3.3: Numerical spectra of A: K = 200, L = 100, J = 10.

    4 Algorithm

    The fact that the coordinates x,y, z of the sources βk,l, k = 1, . . . , K,

    l = 1, . . . , L, form an eigenspace of A (see Section 3.2) enables to com-

    pute them by computing the first three non-trivial eigenvectors ψ1,ψ2,ψ3

    of the sparse matrix A. Taking a sufficiently small J ensures that x,y, z

    appear immediately after ψ0 = 1 in the spectrum of A. However, due to

    the threefold multiplicity of the eigenvalue, the computed eigenvectors may

    be any linear combination of the coordinate vectors. This linear combina-

    tion is easily determined (up to an orthogonal transformation) by using the

    fact that the coordinates must correspond to points on the sphere (i.e., unit

    length vectors). To this end, we look for a 3 × 3 matrix M such that

    X ≡

    − xT −

    − yT −

    − zT −

    = M

    − ψT1 −

    − ψT2 −

    − ψT3 −

    ≡MΨ. (17)

    The diagonal of the KL×KL matrix XTX = ΨTMTMΨ is all ones, be-

    cause the points are on the unit sphere, that is, ‖βk,l‖2 = x2(k, l)+y2(k, l)+

    z2(k, l) = 1. We end up with an overdetermined system of KL linear equa-

    20

  • tions(

    ΨTMTMΨ)

    ii= 1, (18)

    for the 9 entries of MTM . The least squares solution for MTM is then

    followed by an SVD or a Cholesky decomposition to yield M . We can

    recover M only up to an orthogonal transformation O ∈ O(3), because

    MTOTOM = MTM . The reconstruction of the molecule is up to an arbi-

    trary rotation and possibly a reflection (the chirality or handedness cannot

    be determined).

    The locations of the radial lines can be further refined by using the fact

    that same image radial lines correspond to a great circle on S2. In particular,

    such radial lines belong to the same plane (slice). Therefore, we improve the

    estimation of coordinates by using principal component analysis (PCA) for

    groups of L radial lines at a time. Furthermore, we equally space those radial

    lines on the corresponding great circle.

    GCAR is summarized in Algorithm 1. Step 3 of finding pairs of common

    lines can be done efficiently in linear time complexity (instead of comparing

    a quadratic number of pairs) by approximate nearest neighbors algorithms in

    a low dimensional feature space. Note that the algorithm assumes that the

    projection images are centered, for otherwise an arbitrary phase, which alters

    the step of finding common lines, comes into play. Handling non-centered

    projections is addressed in Section 6.1.

    The eigenvector computation is global and takes into account all the

    local pieces of information about common lines. Even if some common lines

    are misidentified, those errors are averaged out in the global eigenvector

    computation. Thus, GCAR should be regarded as a very efficient way of

    integrating the local cryo-EM geometry into a global orientation assignment.

    The construction of the matrix W , as described in Section 3.1, uses all

    pairs of common lines. That is, for each pair of projection images k1 and k2,

    we find the Fourier lines Λk1,l1 and Λk2,l2 such that Λk2,l2 is closest to Λk1,l1 ,

    and use the pair (k1, l1) and (k2, l2) to add edges to the set E in Eqs. 8–9.

    21

  • As can be seen in Fig. 2.2, this corresponds to finding all geodesic circles

    on S2 that pass though βk1,l1. Note however, that the coordinates vectors

    are eigenvectors of A = D−1W even if we use only a few of the geodesic

    circles that go through βk1,l1 . This corresponds to using fewer legs in each

    spider. Moreover, the resulting matrix W is sparser, and so requires less

    memory and its eigenvectors can be computed faster. The key advantage

    of this observation is that we do not need to use all lines determined by

    the(

    K2

    )

    intersection of projection images. We use only pairs of images for

    which Λk1,l1 is very close to Λk2,l2 . This results in fewer misidentifications of

    common-lines, and leads to a more accurate estimation of the orientations.

    This is demonstrated in Section 5.

    Algorithm 1 Outline of GCAR

    Require: Projection images Pk(x, y), k = 1, 2, . . . , K1: Compute the polar Fourier transform P̂k(ρ, ω) (Eq. 4).2: Split each P̂k(ρ, ω) into L radial lines Λk,l (Eq. 5).3: Find common lines Λk1,l1 ≈ Λk2,l2.4: Construct the sparse KL×KL weight matrix W with J ≪ L (following

    Section 3.1).5: Normalize W by its outdegree D and form the averaging operator A =D−1W (Eq. 13).

    6: Compute the first three non-trivial eigenvectors of A: Aψ1 =λψ1, Aψ2 = λψ2, Aψ3 = λψ3.

    7: Unmix x,y, z from ψ1,ψ2,ψ3.8: Refinement: PCA and equally space same image radial lines.

    5 Numerical Examples

    The algorithm was implemented in MATLAB and was tested on two types of

    phantoms. The first phantom consists of a set of ellipsoids, whose projections

    can be computed analytically. This phantom is shown in Fig. 5.1a. The

    second phantom is a 96 × 96 × 96 density map of the E. coli ribosome,

    22

  • whose projections were computed by approximating line integrals through

    the volume. The E. coli phantom is pictured in Fig. 5.5a. All tests used

    K = 100 projection images, with L = 300 radial Fourier lines per projection,

    and 100 samples along each Fourier ray. The projections of the analytic

    phantom were of size 97 × 97 to avoid half pixel shifts involved in even-size

    projections. The projections of the E. Coli phantom were of size 96 × 96

    and center estimation was used to compensate for half pixel shifts. The

    polar Fourier transform was computed as described in [16]. Common-lines

    between two Fourier-transformed projections were found by computing and

    comparing correlations between all Fourier lines of the two projections. No

    knowledge of the orientations or their distribution was used by the algorithm.

    All tests were executed on a quad core Xeon 2.33GHz running Linux.

    Once orientations were determined, the phantom was reconstructed by inter-

    polating theKL Fourier lines into the three-dimensional pseudo-polar grid by

    using nearest-neighbor interpolation followed by an inverse 3D pseudo-polar

    Fourier transform, implemented along the lines of [10, 17].

    Figure 5.1a shows a 3D rendering of the analytic phantom, with several of

    its projections given in Fig. 5.2. The projection orientations for this phantom

    were randomly sampled from the uniform distribution on S2. We computed

    the common line between each pair of projections, that is, the Fourier rays

    Λk1,l1 and Λk2,l2, such that Λk1,l1 is closest to Λk2,l2 . Figure 5.3 shows the

    dissimilarity between Λk1,l1 and Λk2,l2 for each pair of images k1 and k2, sorted

    from the smallest (most similar) to the largest (most different). Such a plot

    allows to pick the threshold for filtering the GCAR matrix, as described in

    Section 4.

    Figure 5.4a shows the spectrum of the operator A. Figure 5.4b presents

    the orientation estimation error. The estimation error for each orientation

    is defined as the angle (in radians) between the true orientation and the

    estimated one. Finally, Fig. 5.1b shows the reconstructed phantom. We

    can see that the reconstructed phantom is related to the original phantom

    23

  • (a) Original (b) Reconstructed

    Figure 5.1: Original and reconstructed analytic phantoms.

    through an orthogonal transformation, which follows from the multiplicity of

    the three-dimensional eigenspace.

    In Figs. 5.5–5.8 we present the results for the E. coli ribosome density

    map. Figure 5.5a shows the reference three-dimensional density map of the

    E. coli ribosome. Several of its projections are given in Fig. 5.6. The orienta-

    tions in which the projections of this phantom were computed are presented

    in Fig. 5.7. The spectrum of the corresponding operatorA is given in Fig. 5.8.

    Note that the multiplicities in this spectrum are different than the multiplic-

    ities of the spectrum in Fig. 5.4a. The second eigenspace in Fig. 5.8 is of

    dimension three as the coordinates x, y, and z are always exact eigenvectors,

    as shown in Section 3.3. However, as opposed to the spectrum in Fig. 5.4a,

    the next eigenspace is not of dimension five. Due to the distribution of the

    projection orientations, shown in Fig. 5.7, the resulting GCAR operator does

    not commute with rotations at all. Hence, the arguments in Section 3.4 do

    not hold and the spectrum in Fig. 5.8 is not the spectrum of the spherical

    harmonics. Finally, two views of the reconstructed density map are given in

    Figs. 5.5b and 5.5c.

    24

  • Figure 5.2: Sample of eight projection images of the analytic phantom.

    0 500 1000 1500 2000 2500 3000 3500 4000 4500 500010

    −7

    10−6

    10−5

    10−4

    10−3

    10−2

    Figure 5.3: Semi-log plot of the dissimilarities between each pair of commonlines.

    25

  • 0 1 2 3 4 5 6 7 8 9 10 11 12 13 140

    0.005

    0.01

    0.015

    0.02

    0.025

    0.03

    0.035

    0.04

    0.045

    0.05

    (a) Spectrum of the GCAR operator|1 − λi|

    0 0.5 1 1.5 2 2.5 3 3.5 4

    x 10−3

    0

    200

    400

    600

    800

    1000

    1200

    1400

    1600

    1800

    (b) Error histogram

    Figure 5.4: Spectrum of the GCAR operator for the analytic phantom andthe estimation error histogram.

    (a) Original

    (b) Reconstructed view 1 (c) Reconstructed view 2

    Figure 5.5: Original and reconstructed E. coli density maps.

    26

  • Figure 5.6: Sample of eight projection images of the E. coli ribosome densitymap.

    −1

    −0.5

    0

    0.5

    1−1

    −0.50

    0.51

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    Figure 5.7: Orientations on S2 used to generate the E. coli projections.

    27

  • 0 1 2 3 4 5 6 7 8 9 10 11 12 13 140

    0.005

    0.01

    0.015

    0.02

    0.025

    0.03

    0.035

    0.04

    0.045

    Figure 5.8: Spectrum of the GCAR operator (|1 − λi|) for the E. coli densitymap.

    6 Extensions

    6.1 Center determination

    The Fourier projection-slice theorem in Section 2.2 requires that for any

    projection P taken in the direction θ, the center of the imaged object is

    projected into the center of the projection image P . Such a center, however,

    has no physical meaning, and any point in the three-dimensional object space

    can be chosen as the center, by simply putting the origin of the coordinate

    system at that point. Equation 3 states that this point should be projected

    into the origin of the coordinate system in all projections.

    In practice, each projection image is segmented from a much larger micro-

    graph, containing many projection images, by roughly estimating the bound-

    ing box of the imaged molecule’s copy. The projections obtained by such a

    segmentation procedure would not satisfy Eq. 3 simultaneously, because their

    centers are inconsistent. This means, for example, that defining the center

    of each projection as its center of mass would lead to inconsistencies.

    As a consequence, the input to the cryo-EM problem is a set of projections

    Qg1, . . . , QgK , where each projection Qgk contains some unknown shift with

    respect to its unshifted version Pgk to which the common line property of

    Eq. 3 can be applied; that is, Qgk(xk, yk) = Pgk(x

    k + ∆xk, yk + ∆yk), where

    28

  • ∆xk and ∆yk are the unknown shifts.

    The 2D Fourier transform (see also Eq. 4) is given

    f̂(ωx, ωy) =1

    2π2

    ∫∫ ∞

    −∞

    f(x, y)e−ı(xωx+yωy)dx dy.

    By the Fourier shift theorem, if g(x, y) = f(x + ∆x, y + ∆y) for some fixed

    shifts ∆x and ∆y, then,

    ĝ(ωx, ωy) = f̂(ωx, ωy)eı(∆xωx+∆yωy). (19)

    Let Pg1 and Pg2 be two unshifted projections (centered with respect to the

    center of the underlying three dimensional object). We assume that Pg1 uses

    the coordinate system (x1, y1), and Pg2 uses the coordinate system (x2, y2).

    Let

    Qg1(x1, y1) = Pg1(x

    1 + ∆x1, y1 + ∆y1),

    Qg2(x2, y2) = Pg2(x

    2 + ∆x2, y2 + ∆y2)

    be the translated versions of Pg1 and Pg2, shifted by (∆x1,∆y1) and (∆x2,∆y2),

    respectively. The Fourier shift theorem (Eq. 19) implies

    Q̂g1(ω1x, ω

    1y) = P̂g1(ω

    1x, ω

    1y)e

    ı(∆x1ω1x+∆y1ω1y),

    Q̂g2(ω2x, ω

    2y) = P̂g2(ω

    2x, ω

    2y)e

    ı(∆x2ω2x+∆y2ω2y).

    (20)

    Suppose that the common line of the projections P̂g1 and P̂g2 is (r cos θ1, r sin θ1)

    in P̂g1 and (r cos θ2, r sin θ2) in P̂g2, with θ

    1 and θ2 measured from the wx-axis

    in P̂g1 and P̂g2, respectively. Along the common line

    P̂g1(r cos θ1, r sin θ1) = P̂g2(r cos θ

    2, r sin θ2), (21)

    29

  • and so,

    Q̂g1(r cos θ1, r sin θ1)e−ır(∆x

    1 cos θ1+∆y1 sin θ1)

    = Q̂g2(r cos θ2, r sin θ2)e−ır(∆x

    2 cos θ2+∆y2 sin θ2),

    from which we get

    1

    rarg

    Q̂g1(r cos θ1, r sin θ1)

    Q̂g2(r cos θ2, r sin θ2)

    = µg1,g2 (22)

    with

    µg1,g2 = ∆x1 cos θ1 + ∆y1 sin θ1 − ∆x2 cos θ2 − ∆y2 sin θ2. (23)

    Given K projection images, there are 2K unknowns(

    ∆xk,∆yk)

    and(

    K2

    )

    equations of the form of Eq. 22. Thus we form the(

    K2

    )

    × 2K matrix system

    of linear equations given by Eq. 22, and solve it using least squares. Note

    that this linear system is very sparse as each row contains only four nonzero

    elements. The resulting matrix has a null-space of dimension three, which

    reflects the fact that arbitrarily moving the origin of the object space R3

    induces another set of consistent translations(

    ∆xk,∆yk)

    in the projections,

    which also satisfy Eq. 22. As in the case of constructing the matrix W

    in Section 3.1, we need not use all(

    K2

    )

    equations, but only equations that

    correspond to highly similar common lines. We can also further filter the

    system by choosing only equations that correspond to pairs for which the

    left hand side in Eq. 22 in nearly constant for various values of r. Although

    in theory the left hand side of Eq. 22 should be constant for all r, this is not

    the case in practice due to discretization, noise, and measurement errors.

    To form the translation estimation equations described above, we need to

    detect common lines between pairs of projections in the presence of unknown

    relative shifts. As a result of Eq. 21

    |P̂g1(r cos θ1, r sin θ1)| = |P̂g2(r cos θ

    2, r sin θ2)|,

    30

  • and from Eq. 20 we get

    |Q̂g1(r cos θ1, r sin θ1)| = |Q̂g2(r cos θ

    2, r sin θ2)|.

    Hence, to detect common lines between projections that were shifted by some

    unknown shift, we take the polar Fourier transform of each projection, and

    find common lines between the magnitude of the Fourier rays.

    It may be that two Fourier rays have the same absolute values, although

    these are not the common line between the two projections. To overcome this

    problem we find not only the closest pair of rays in |Q̂g1 | and |Q̂g2|, but several

    such pairs. For each pair we assume it is the common line and estimate the

    phase factor µg1,g2 in Eq. 23 using Eq. 22. Thus we can compute the similarly

    between the rays P̂g1(r cos θ1, r sin θ1) and P̂g2(r cos θ

    2, r sin θ2) by computing

    the correlation between Q̂g1(r cos θ1, r sin θ1)e−ırµg1,g2 and Q̂g2(r cos θ

    2, r sin θ2),

    and choosing the pair θ1 and θ2 that brings this correlation to maximum. We

    then use these θ1 and θ2 to form a common line equation of the form of Eq. 22.

    This results, as explained above, in(

    K2

    )

    equations for the 2K unknown shifts,

    which we solve using least-squares.

    The performance of the center determination algorithm is demonstrated

    in Fig. 6.1. We used K = 100 projections of the analytic phantom (Fig. 5.1a),

    each of size 97× 97 pixels. An odd size was chosen for the projection images

    to avoid inherent half-pixel shifts associated with even sampling sizes. Each

    projection was randomly shifted in the x and y directions by some random

    shift of up ±20 pixels in each direction. For each projection we computed L =

    300 radial Fourier lines, each with n = 100 samples in the radial direction.

    Figure 6.1a shows the 10 smallest singular values of the center estimation

    system, given by Eq. 22. The null space of dimension three is apparent.

    Figure 6.1b shows the shifts estimation error. Each bar corresponds to the

    absolute estimation error (in pixels) in one of the K = 100 projections.

    31

  • 1 2 3 4 5 6 7 8 9 100

    0.5

    1

    1.5

    2

    2.5

    3

    (a) Smallest singular values of the(

    K

    2

    )

    ×2K sparse matrix obtained from Eq. 22

    10 20 30 40 50 60 70 80 90 1000

    0.002

    0.004

    0.006

    0.008

    0.01

    0.012

    0.014

    0.016

    0.018

    0.02

    (b) Absolute estimation error

    Figure 6.1: Performance of the center estimation algorithm.

    (a) Molecule of type 1 (b) Molecule of type 2

    Figure 6.2: Two types of molecules (3D rendering of the analytic phantoms).

    6.2 Heterogeneity problem

    Suppose our data set contains projections of two different types of molecules.

    In this section we describe how to use the spectral properties of the operator

    A defined in Section 3.2 to discriminate between the two different molecules.

    We will accompany the explanation with an example using two analytic phan-

    toms, depicted in Figs. 6.2a and 6.2b. The example shows only two types

    of molecules, but essentially the same algorithm can be used to handle more

    than two types.

    Suppose we have a mix of K1 projections from the first molecule and K2

    32

  • projections from the second molecule. Assume for the sake of presentation

    that the first K1 projections belong to type 1, and the next K2 projections

    belong to type 2. Suppose that we find the common line between any pair of

    images. For each of the(

    K1+K22

    )

    pairs of images, we obtain the lines Λk1,l1

    and Λk2,l2, which are supposed to be a common line Λk1,l1 = Λk2,l2 . We

    measure the dissimilarity between Λk1,l1 and Λk2,l2 and obtain a dissimilarity

    coefficient disk1,k2 corresponding to the pair of projections (k1, k2). We expect

    that on average, the dissimilarity coefficient for two projections that belong

    to the same type of molecule will be smaller (more similar) than for two

    projections of different types. Thus, if we sort the dissimilarity coefficients

    dissk1,k2, k1, k2 = 1, . . . , K1 + K2 from smallest to largest, we expect that

    the first part of the sorted array will correspond to dissimilarities between

    projections from the same type of molecule, and the second part of the sorted

    array, which corresponds to larger dissimilarity coefficients, will correspond

    to common lines between projections of different types.

    To demonstrate this point, we generated a mix that contains K1 = 100

    random projections of the molecule of type 1 (Fig. 6.2a) and K2 = 100

    random projections of the molecule of type 2 (Fig. 6.2b). The random ori-

    entations used for each type of molecule are independent. Each projection

    is of size 96 × 96 pixels. To determine common lines, we computed L = 300

    radial Fourier line in each projection (angular resolution), and used n = 100

    samples on each Fourier line (radial resolution). We computed the dissimi-

    larity coefficient for each pair of images. The sorted dissimilarity coefficients

    are shown in Fig. 6.3. As expected, the first part of the sorted list contains

    small dissimilarity coefficients (roughly half of the list since K1 = K2 = 100).

    The second part of the sorted list contains significantly larger values.

    If we filter the GCAR matrix W , defined in Section 3.1, as descried in

    Section 4, by retaining only common lines that correspond to the first part of

    the graph in Fig. 6.3, then the resulting matrix of size (K1+K2)L×(K1+K2)L

    is a block matrix with two non-interacting blocks. Hence, the multiplicities

    33

  • 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

    x 104

    10−7

    10−6

    10−5

    10−4

    10−3

    10−2

    Figure 6.3: Dissimilarity coefficient sorted from smallest to largest.

    of the eigenvalues will be doubled compared to the case of a single type of

    molecule. The spectrum of A for the given heterogeneity example is given

    in Fig. 6.4.

    The first eigenspace (corresponding to eigenvalue 1, or |1 − λ1| = 0 in

    Fig. 6.4) is of dimension two, and the second eigenspace is of dimension

    six. The first eigenvector, which in the single molecule case is the all-ones

    vector, now becomes piecewise constant. In fact, by orthogonalizing the first

    eigenspace of dimension two relative to the all-ones vector, we get a piecewise

    constant vector w ∈ R(K1+K2)L such that wj > 0 if Fourier line j belongs to

    a projection of the first type, and wj < 0 otherwise. The first eigenvector of

    the mixed GCAR matrix is presented in Fig. 6.5. Figure 6.5 corresponds to

    the case where the K1 +K2 projections are randomly shuffled. Only 15000

    out of the 60000 entries of the first eigenvector are shown.

    We can therefore partition all Fourier lines according to the type of their

    underlying molecule, and thus classify the projections according to their type.

    Once partitioned into two classes, the orientations in each class can be deter-

    mined separately. Figures 6.6a and 6.6b show the sorted dissimilarity coeffi-

    34

  • 0 1 2 3 4 5 6 7 8 90

    0.005

    0.01

    0.015

    0.02

    0.025

    Figure 6.4: Spectrum of the filtered GCAR matrix (|1 − λi|) constructedfrom a mix of two types of molecules.

    0 5000 10000 15000−5

    −4

    −3

    −2

    −1

    0

    1

    2

    3

    4

    5x 10

    −3

    Figure 6.5: First few values of the first eigenvector of the mixed GCARmatrix (after orthogonalizing to the all-ones vector).

    35

  • 0 500 1000 1500 2000 2500 3000 3500 4000 4500 50000

    0.2

    0.4

    0.6

    0.8

    1

    1.2x 10

    −4

    (a) Molecule of type 1

    0 500 1000 1500 2000 2500 3000 3500 4000 4500 50000

    1

    x 10−4

    (b) Molecule of type 2

    Figure 6.6: Dissimilarity coefficients for each type of molecule.

    0 1 2 3 4 5 6 7 8 90

    0.005

    0.01

    0.015

    0.02

    0.025

    0.03

    0.035

    0.04

    (a) Molecule of type 1

    0 1 2 3 4 5 6 7 8 90

    0.005

    0.01

    0.015

    0.02

    0.025

    0.03

    0.035

    0.04

    0.045

    (b) Molecule of type 2

    Figure 6.7: Spectrum of the GCAR matrix for each type of molecule.

    cient for each type of molecule after classification. The plots in Figs. 6.6a and

    6.6b are then used to separately filter the two GCAR matrices corresponding

    to the individual molecules.

    Figures 6.7a and 6.7b show the spectrum of the GCAR matrix of each

    class. The spectrum for each class is as predicted: the first eigenvalue is 1, and

    the next subspace is of dimension three. We then estimate the orientations of

    the projections in each class separately. Histograms of the error estimation

    are given in Figs. 6.8a and 6.8b.

    The advantage of the above procedure is that it uses all the data simul-

    36

  • 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.0450

    500

    1000

    1500

    2000

    2500

    3000

    estimation error

    (a) Molecule of type 1

    0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.040

    1000

    2000

    3000

    4000

    5000

    6000

    estimation error

    (b) Molecule of type 2

    Figure 6.8: Histogram of the orientation estimation error for each type ofmolecule.

    taneously to separate the classes, as opposed to other methods, which try to

    classify each projection separately.

    6.3 Particle selection

    The data acquisition process in cryo-electron microscopy generates many

    projection images which are corrupted by noise, improper segmentation, and

    overlapping particles, to name but a few. It is desirable to be able to remove

    such particles from the data set before any reconstruction process takes place.

    The matrixW , defined in Section 3.1, can be used to identify such corrupted

    projections. Once constructing the matrix W , we filter it as described in

    Section 4. At this point we inspect how many projections intersect a given

    projection such that the common line between the two projections is highly

    similar. In theory, in a noiseless setting, any two projections should have a

    common line. That is, each projection image should have a total of O(K)

    interactions with other projections. When a projection is corrupted by noise

    or is otherwise inconsistent with the rest of the projections, there would

    be only a few such intersections. This would suggest that this projection

    is inconsistent with the rest of the data set and should be rejected. This

    37

  • Figure 6.9: Corrupted projections.

    again demonstrates the principle that all the data at once should be used to

    determine which projections are to be used for reconstruction: a projection

    is “good” if many other projections “say” it is good, by having a common

    line with it.

    To demonstrate this idea we generated K = 100 projections of the an-

    alytic phantom in Fig. 5.1a, and corrupted 10 projections by a rectangle of

    Gaussian noise of size up to 10×10 pixels. See Fig. 6.9 for several of these cor-

    rupted projections. We then constructed the matrix W as described above,

    filtered it, and rejected projections that intersect with less than 10 percent

    of the other projections. Figure 7.1 shows the number of intersections for

    each projection. The first 20 projections are the corrupted ones. It is clear

    that by removing projections with a few intersections we retain the “good”

    projections.

    7 Summary and Future Work

    In this technical report we introduced a new methodology for the three di-

    mensional cryo-EM structure determination. Our GCAR algorithm incorpo-

    rates the Fourier-projection slice theorem into a novel construction of a graph,

    followed by an efficient calculation of a few eigenvectors of the normalized

    sparse adjacency matrix. The resulting eigenvectors reveal the projection ori-

    entations in a globally consistent manner. We demonstrated the success of

    the method when applied to artificially produced projection images, as well

    38

  • as its applicability to the image centering and the heterogeneity problems.

    The reader must be asking herself if this method would also be successful

    in practice, when faced with real noisy images, rather than artificially pro-

    duced clean images. We have good reasons to believe that this would be the

    case, but at this point of time, we prefer leaving speculation aside. One has

    to keep in mind that this technical report summarizes a work which is still in

    progress, and we hope that soon enough we will be able to provide a definite

    answer.

    Not included in this technical report are preliminary results regarding

    the behavior and success of the GCAR algorithm when the input images are

    corrupted by white Gaussian noise. Detection of common lines in the pres-

    ence of noise is much more difficult: out of all detected common lines, only

    a small percentage are actually true common lines. Although the resulting

    embedding found by the GCAR algorithm is distorted and cannot be used

    to reveal the orientations, it can be iteratively improved until convergence

    to a globally consistent embedding is obtained. In each iteration, we ignore

    common lines that do not agree with the previously obtained embedding.

    This iterative procedure, which resembles well known procedures in robust

    estimation, such as the iterative weighted least squares procedure, cleans up

    the noisy graph and enables successful reconstructions.

    Clearly, denoising of either projection images or radial Fourier lines is a

    major theme of our future work. We plan to apply both classical and modern

    denoising methods from image and signal processing. There are two main

    approaches for denoising the projection images. In the first approach, each

    projection image will be denoised separately, and the denoised images will

    be compared to find the common lines. In the second approach, we will

    try to denoise many images at once, since similar features should appear

    in different images. A complete discussion of denoising methods and their

    practical success will be the subject of a future report.

    39

  • 10 20 30 40 50 60 70 80 90 1000

    10

    20

    30

    40

    50

    60

    70

    80

    90

    Figure 7.1: Number of intersections of each projection with other projectionsafter filtering the GCAR matrix.

    References

    [1] J. Frank. Three-Dimensional Electron Microscopy of Macromolecular

    Assemblies: Visualization of Biological Molecules in Their Native State.

    Oxford, 2006.

    [2] L. Wang and F. J. Sigworth. Cryo-em and single particles. Physiology

    (Bethesda), 21:13–8, 2006. Review. PMID: 16443818 [PubMed - indexed

    for MEDLINE].

    [3] R. Henderson. Realizing the potential of electron cryo-microscopy. Q

    Rev Biophys, 37(1):3–13, 2004. Review. PMID: 17390603 [PubMed -

    indexed for MEDLINE].

    [4] W. Chiu, M. L. Baker, W. Jiang, M. Dougherty, and M .F. Schmid.

    Electron cryomicroscopy of biological machines at subnanometer resolu-

    tion. Structure, 13(3):363–372, 2005. Review. PMID: 15766537 [PubMed

    - indexed for MEDLINE].

    [5] M. Van Heel. Angular reconstitution: a posteriori assignment of projec-

    tion directions for 3D reconstruction. Ultramicroscopy, 21(2):111–123,

    1987. PMID: 12425301 [PubMed - indexed for MEDLINE].

    40

  • [6] A. E. Leschziner and E. Nogales. Visualizing flexibility at molecular res-

    olution: analysis of heterogeneity in single-particle electron microscopy

    reconstructions. Annu Rev Biophys Biomol Struct, 36:43–62, 2007. Re-

    view. PMID: 17201674 [PubMed - indexed for MEDLINE].

    [7] S. H. Scheres, H. Gao, M. Valle, G. T. Herman, P. P. Eggermont,

    J. Frank, and J. M. Carazo. Disentangling conformational states of

    macromolecules in 3D-em through likelihood optimization. Nature Meth-

    ods, 4(1):27–29, 2007. Epub 2006 Dec 10.

    [8] F. Natterer. The Mathematics of Computerized Tomography. Classics in

    Applied Mathematics. SIAM: Society for Industrial and Applied Math-

    ematics, 2001.

    [9] A. Dutt and V. Rokhlin. Fast Fourier transforms for nonequispaced

    data. SIAM Journal on Scientific Computing, 14(6):1368–1393, 1993.

    [10] A. Averbuch and Y. Shkolnisky. 3D Fourier based discrete radon trans-

    form. Applied and Computational Harmonic Analysis, 15(1):33–69,

    2003.

    [11] R. R. Coifman, Y. Shkolnisky, F. J. Sigworth, and A. Singer. Graph

    laplacian tomography from unknown random projection. IEEE Trans-

    actions on Image Processing, To appear.

    [12] S. T. Roweis and L. K. Saul. Nonlinear dimensionality reduction by

    locally linear embedding. Science, 290(5500):2323–2326, 2000.

    [13] D. L. Donoho and C. Grimes. Hessian eigenmaps: Locally linear embed-

    ding techniques for high-dimensional data. Proceedings of the National

    Academy of Sciences, 100(10):5591–5596, 2003.

    [14] M. Belkin and P. Niyogi. Laplacian eigenmaps for dimensionality re-

    duction and data representation. Neural Computation, 15:1373–1396,

    2003.

    41

  • [15] R. R. Coifman, S. Lafon, A. B. Lee, M. Maggioni, B. Nadler, F. Warner,

    and S. W. Zucker. Geometric diffusions as a tool for harmonic analysis

    and structure definition of data: Diffusion maps. Proceedings of the

    National Academy of Sciences, 102(21):7426–7431, 2005.

    [16] L. Greengard and J.-Y. Lee. Accelerating the nonuniform fast Fourier

    transform. SIAM Review, 46(3):443–454, 2004.

    [17] A. Averbuch, R. R. Coifman, D. L. Donoho, M. Israeli, and Y. Shkol-

    nisky. A framework for discrete integral transformations I – the pseudo-

    polar Fourier transform. SIAM Journal on Scientific Computing, To

    appear.

    42


Recommended